Sie sind auf Seite 1von 41

TREATISE OF PLANE GEOMETRY

THROUGH GEOMETRIC ALGEBRA

Ramon Gonzlez Calvet

TIMSAC No. 1

CONTENTS
First Part: The Euclidean vector plane and complex numbers
1. Euclidean vectors and their operations
Vector addition, 1.- Product of a vector and a real number, 2.- Product of two vectors,
2.- Product of three vectors: associative property, 5. Product of four vectors, 7.- Inverse
and quotient of two vectors, 7.- Priority of algebraic operations, 8.- Geometric algebra
of the vector plane, 9.- Exercises, 9.
2. A vector basis for the Euclidean plane
Linear combination of two vectors, 10.- Basis and components, 10.- Orthonormal bases,
11.- Applications of formulae for products, 11.- Exercises, 12.
3. Complex numbers
The subalgebra of complex numbers, 13.- Binomial, polar and trigonometric form of a
complex number, 13.- Algebraic operations with complex numbers, 14.- Permutation of
complex numbers and vectors, 17.- The complex plane, 18.- Complex analytic
functions, 19.- Fundamental theorem of algebra, 24.- Exercises, 26.
4. Transformations of vectors
Rotations, 27.- Axial symmetries, 28.- Inversions, 29.- Dilations, 30.- Exercises, 30
Second Part: Geometry of the Euclidean plane
5. Points and straight lines
Translations, 31.- Coordinate systems, 31.- Barycentric coordinates, 33.- Distance
between two points and area, 33.- Condition of collinearity of three points, 35.Cartesian coordinates, 36.- Vectorial and parametric equations of a line, 36.- Algebraic
equation and distance from a point to a line, 37.- Slope and intercept equations of a line,
40.- Polar equation of a line, 41.- Intersection of two lines and pencil of lines, 41.- Dual
coordinates, 43.- Desarguess theorem, 48.- Exercises, 50.
6. Angles and elemental trigonometry
Sum of the angles of a polygon, 53.- Definition of trigonometric functions and
fundamental identities, 54.- Angle inscribed in a circle and double-angle identities, 55.Addition of vectors and sum of trigonometric functions, 56.- Product of vectors and
addition identities, 57.- Rotations and de Moivre's identity, 58.- Inverse trigonometric
functions, 59.- Exercises, 60.
7. Similarities and simple ratio
Direct similarity (similitude), 61.- Opposite similarity, 62.- Menelaus theorem, 63.Cevas theorem, 64.- Homothety and simple ratio, 65.- Exercises, 67.
8. Properties of triangles
Area of a triangle, 68.- Medians and centroid, 69.- Perpendicular bisectors and
circumcentre, 70.- Angle bisectors and incentre, 72.- Altitudes and orthocentre, 73.Euler's line, 76.- Fermat's theorem, 77.- Exercises, 78.
XII

9. Circles
Algebraic and Cartesian equations, 80.- Intersections of a line with a circle, 80.- Power
of a point with respect to a circle, 82.- Polar equation, 82.- Inversion with respect to a
circle, 83.- The nine-point circle, 85.- Cyclic and circumscribed quadrilaterals, 87.Angle between circles, 89.- Radical axis of two circles, 89.- Exercises, 91.
10. Cross ratios and related transformations
Complex cross ratio, 92.- Harmonic characteristic and ranges, 94.- Homography
(Mbius transformation), 96.- Projective cross ratio, 99.- Points at infinity and
homogeneous coordinates, 102.- Perspectivity and projectivity, 103.- Projectivity as a
tool for theorem demonstrations, 108.- Homology, 110.- Exercises, 115.
11. Conics
Conic sections, 117.- Two foci and two directrices, 120.- Vectorial equation, 121.Chasles' theorem, 122.- Tangent and perpendicular to a conic, 124.- Central equations
for ellipse and hyperbola, 126.- Diameters and Apollonius' theorem, 128.- Conic
passing through five points, 131.- Pencil of conics passing through four points, 133.Conic equation in barycentric coordinates and dual conic, 133.- Polarities, 135.Reduction of the conic matrix to diagonal form, 136.- Exercises, 137.
Third part: Pseudo-Euclidean geometry
12. Matrix representation and hyperbolic numbers
Rotations and the representation of complex numbers, 139.- The subalgebra of
hyperbolic numbers, 140.- Hyperbolic trigonometry, 141.- Hyperbolic exponential and
logarithm, 143.- Polar form, powers and roots of hyperbolic numbers, 144.- Hyperbolic
analytic functions, 147.- Analyticity and square of convergence of power series, 150.About the isomorphism of Clifford algebras, 152.- Exercises, 153.
13. The hyperbolic or pseudo-Euclidean plane
Hyperbolic vectors, 154.- Inner and outer products of hyperbolic vectors, 155.- Angles
between hyperbolic vectors, 156.- Congruence of segments and angles, 158.Isometries, 158.- Theorems about angles, 160.- Distance between points, 160.- Area in
the hyperbolic plane, 161.- Diameters of the hyperbola and Apollonius theorem, 163.The law of sines and cosines, 164.- Hyperbolic similarity, 167.- Power of a point with
respect to a hyperbola with constant radius, 168.- Exercises, 169.
Fourth part: Plane projections of three-dimensional spaces
14. Spherical geometry in the Euclidean space
The geometric algebra of the Euclidean space, 170.- Spherical trigonometry, 172.- The
dual spherical triangle of a given triangle, 175.- Right spherical triangles and Napiers
rule, 176.- Area of a spherical triangle, 176.- Properties of the projections of the
spherical surface, 177.- Central or gnomonic projection, 177.- Stereographic projection,
180.- Orthographic projection, 181.- Lamberts azimuthal equivalent projection, 182.Spherical coordinates and cylindrical equidistant (plate carr) projection, 183.- Mercator

XIII

projection, 184.- Cylindrical equivalent projection, 184.- Conic projections, 185.Exercises, 186.
15. Hyperboloidal geometry in the pseudo-Euclidean space
The geometric algebra of the pseudo-Euclidean space, 189.- The hyperboloid of two
sheets (Lobachevskian surface), 191.- Central projection (Beltrami disk), 192.Lobachevskian trigonometry, 197.- Stereographic projection (Poincar disk), 199.Azimuthal equivalent projection, 201.- Weierstrass coordinates and cylindrical
equidistant projection, 202.- Cylindrical conformal projection, 203.- Cylindrical
equivalent projection, 204.- Conic projections, 204.- About the congruence of geodesic
triangles, 206.- The hyperboloid of one sheet, 206.- Central projection and arc length on
the one-sheeted hyperboloid, 207.- Cylindrical projections, 208.- Cylindrical central
projection, 209.- Cylindrical equidistant projection, 210.- Cylindrical equivalent
projection, 210.- Cylindrical conformal projection, 210.- Area of a triangle on the onesheeted hyperboloid, 211.- Trigonometry of right triangles, 214.- Hyperboloidal
trigonometry, 215.- Dual triangles, 218.- Summary, 221.- Comment about the names of
the non-Euclidean geometry, 222.- Exercises, 222.
16. Solutions to the proposed exercises
1. Euclidean vectors and their operations, 224.- 2. A vector basis for the Euclidean
plane, 225.- 3. Complex numbers, 227.- 4. Transformations of vectors, 230.- 5. Points
and straight lines, 231.- 6. Angles and elemental trigonometry, 241.- 7. Similarities and
simple ratio, 244.- 8. Properties of triangles, 246.- 9. Circles, 257.- 10. Cross ratios and
related transformations, 262.- 11. Conics, 266.- 12. Matrix representation and
hyperbolic numbers, 273.- 13. The hyperbolic or pseudo-Euclidean plane, 275.- 14.
Spherical geometry in the Euclidean space, 278.- 15. Hyperboloidal geometry in the
pseudo-Euclidean space, 284.
Bibliography, 293.
Internet bibliography, 296.
Index, 298.
Chronology of the geometric algebra, 305.

XIV

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

FIRST PART: THE EUCLIDEAN VECTOR PLANE AND


COMPLEX NUMBERS
Points and vectors are the main elements of plane geometry. A point is
conceived (but not defined) as a geometric element without extension, infinitely small,
that has position and is located at a certain place in the plane. A vector is defined as an
oriented segment, that is, a piece of a straight line having length and direction. A vector
has no position and can be translated anywhere. It is usually called a free vector. If we
place the end of a vector at a point, then its head determines another point so that a
vector represents the translation from the first point to the second.
Taking into account the distinction between points and vectors, the portion of the
book devoted to the plane Euclidean geometry has been divided into two parts. In the
first one, vectors and their algebraic properties are studied, which is enough for many
scientific and engineering branches. In the second part, points are introduced and the
affine geometry is then studied.
All elements in geometric algebra (scalars, vectors, bivectors and complex
numbers) are denoted by lowercase Latin characters and angles with Greek characters.
Capital Latin characters will usually denote points in the plane. As you will see, the
geometric product is not commutative so that fractions can only be written for real and
complex numbers. Since the geometric product is associative, the inverse of a certain
element on the left and on the right is the same, that is, there is a unique inverse for each
element of the geometric algebra, which is indicated by the superscript 1. Moreover,
due to the associative property, all factors in a product are written without parentheses.
In order to make the reading easy, neither theorems nor corollaries nor equations have
been numerated. When a definition is introduced, the definite element is marked with
italic characters that catch the readers attention and help to find the definition once
again.
1. EUCLIDEAN VECTORS AND THEIR OPERATIONS
A vector is an oriented segment, having length and direction but no position, that
is, it can be placed anywhere without changing its orientation. Vectors can represent
many physical magnitudes such as force, velociity, and also geometric magnitudes such
as translation.
We define two algebraic operations, the addition and the product of vectors,
which generalise addition and multiplication of real numbers.
Vector addition
The addition of two vectors u + v
is defined as the vector going from the
end of the vector u to the head of v when
the head of u makes contact with the end
of v (upper triangle in figure 1.1).
Making the construction for v + u, that is,
placing the end of u at the head of v
(lower triangle in figure 1.1), we can see
that the addition vector is the same.

Figure 1.1

RAMON GONZALEZ CALVET

Therefore, the vector addition has the commutative property:


u+v=v+u
and the parallelogram rule follows: the addition of two vectors is the diagonal of the
parallelogram formed by both vectors.
The associative property is the result of
this definition because (u+v)+w or
u+(v+w) is the vector closing the
polygon formed by the three vectors, as
shown in figure 1.2.
The neutral element of the
vector addition is the null vector, which
has no length. Hence the opposite
vector of u is defined as the vector u
with the same orientation but opposite
Figure 1.2
direction, which added to the initial
vector gives the null vector:
u + ( u) = 0
Product of a vector and a real number
The product of a vector and a real number (or scalar) k is defined as a vector
with the same direction but whose length has been increased k times (figure 1.3). If the
real number is negative, then the direction is opposite. The geometric definition implies
the commutative property:
ku=uk

Figure 1.3

Two vectors u, v with the same


direction are proportional because there
is always a real number k such that v =
k u , that is, k is the quotient of both
vectors:
k = u 1 v = v u 1
Two vectors with different
directions are said to be linearly
independent.
Product of two vectors
The product of two vectors will be called the geometric product in order to be
distinguished from other vector products currently used. Nevertheless, I hope that these
other products will play a secondary role when the geometric product becomes the most

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

used, a near event this book will forward. The adjective geometric will then not be
necessary.
We want the geometric product of two vectors to have the following properties:
1) To be distributive over vector addition:
u(v+w)=uv+uw
2) The square of a vector must be equal to the square of its length. By
definition, the length (or norm) of a vector is a positive number and it is
denoted by | u |:
u2 = | u |2
3) The mixed associative property must exist between the product of vectors
and the product of a vector and a real number.
k(uv)=(ku)v= kuv
k(lu)=(kl)u= klu
where k, l are real numbers and u, v vectors. Therefore, parentheses are not
needed in this case.
These properties allow us to deduce the product. Let us suppose that c is the
addition of two vectors a, b and let us calculate its square applying the distributive
property:
c=a+b
c2 = ( a + b )2 = ( a + b ) ( a + b ) = a2 + a b + b a + b2
We have to preserve the order of the factors because we do not know whether the
product is commutative or not.
If a and b are orthogonal vectors, the Pythagorean theorem applies and then:
ab

c2 = a2 + b2

ab+ba=0

ab=ba

That is, the product of two perpendicular vectors is anticommutative.


If a and b are proportional vectors then:
a || b

b = k a, k real

ab=aka=kaa=ba

because of the commutative and mixed associative properties of the product of a vector
and a real number. Therefore the product of two proportional vectors is commutative. If
c is the addition of two vectors a, b with the same direction and sense, we have:
|c|=|a|+|b|
c2 = a2 + b2 + 2 | a | | b |

RAMON GONZALEZ CALVET

ab=|a||b|

(a, b) = 0

But if these vectors have opposite directions:


|c|=|a||b|
c2 = a2 + b2 2 | a | | b |
ab=|a||b|

(a, b) =

What is the product of two vectors with any directions? Due to the distributive
property, the product is resolved into a product by the proportional component b|| and
another by the orthogonal component b:
a b = a ( b|| + b ) = a b|| + a b
The product of one vector by the proportional component of the other one is
called the inner product (also scalar product) and denoted by a dot (figure 1.4).
Taking into account that the projection of b onto a is proportional to the cosine of the
angle between both vectors, one finds:
a b = a b|| = | a | | b | cos
Figure 1.4
The inner product is always a real
number. For example, the work made by
some force acting on a body is the inner
product of the force and the walked path.
Since the commutative property has been
deduced for the product of vectors with the
same direction, the inner product is also
commutative:
ab=ba
The product of one vector by the orthogonal component of the other vector is
called the outer product (also exterior product) and it is denoted by the symbol :
a b = a b
The outer product represents the area of the
parallelogram formed by both vectors (figure 1.5):
a b = a b = absin
Since the outer product is a product of orthogonal
vectors, it is anticommutative:
ab=ba

Figure 1.5

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

Some examples of physical magnitudes that are outer products are angular
momentum, torque and magnetic field.
When two vectors are permuted, the oriented angle is reversed. Then, its cosine
remains unchanged while its sine changes the sign. Thus, the inner product is
commutative while the outer product is anticommutative. Now we can rewrite the
geometric product as the sum of both products:
ab=ab+ab
From here, the inner and outer
products can be written using the geometric
product:
ab =

Figure 1.6

ab+ba
2

a b=

abba
2

In conclusion, the geometric product of two proportional vectors is commutative


whereas the product of two orthogonal vectors is anticommutative, just for the pure
cases of outer and inner products. The outer, inner and geometric products of two
vectors only depend upon the norms of both vectors and the angle between them. When
both vectors are rotated preserving the angle they form, all three products are also
preserved (figure 1.6).
What is the absolute value of the product of two vectors? Since the inner and
outer products are linearly independent and orthogonal magnitudes, the norm of the
geometric product must be calculated by means of a generalisation of the Pythagorean
theorem:
ab=ab+ab

a b 2 = a b2 +a b2

a b 2 = a 2 b 2 ( cos2 + sin2 ) = a 2 b 2
That is, the norm of the geometric product is the product of the norms of each
vector:
a b = a b
Product of three vectors: associative property
It is demanded as the fourth property that the product of three vectors should be
associative:
4)

u(vw)=(uv)w=uvw

Hence we can remove parentheses in multiple products, and with the foregoing
properties we can deduce how the product operates upon vectors.

RAMON GONZALEZ CALVET

We want to multiply a vector a by a product b c of two vectors. We ignore the


result of the product of three vectors with different orientations except when two
adjacent factors are proportional. We have seen that the product of two vectors only
depends on the angle between them. Therefore the parallelogram formed by b and c can
be rotated until b has, in the new orientation, the same direction as a. If b' and c' are the
vectors b and c with the new orientation (figure 1.7) then:
b c = b' c'
a ( b c ) = a ( b' c' )

Figure 1.7

and by the associative property:


a ( b c ) = ( a b' ) c'
Since a and b' have the same
direction, a b' = ab is a real
number and the triple product is a
vector with the direction of c' whose
length is increased by this amount:
a ( b c ) = ab c'
Consequently, the norm of the product of three vectors is the product of their
norms:
a b c = abc
On the other hand, we can first
multiply a by b, and then we can rotate
the parallelogram formed by both
vectors until b has, in the new
orientation, the same direction as c
(figure 1.8). As a result:

Figure 1.8

( a b ) c = a'' ( b'' c ) = a'' bc


Although
this
geometric
construction differs from the foregoing
one, the figures clearly show that the
triple product yields the same vector, as expected from the associative property.
Moreover, we have:
( a b ) c = a'' bc = cb a'' = c b'' a'' = c ( b a )
That is, the triple product fulfils the permutative property:
abc=cba

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

Every vector can be permuted with a vector located two positions farther in a product,
although it does not commute with the neighbouring vectors. The permutative property
implies that any pair of vectors in a product separated by an odd number of vectors can
be permuted. For example:
abcd=adcb=cdab=cbad
The permutative property is characteristic of the plane and it is also valid for the
space whenever the three vectors are coplanar. This property is related to the fact that
the product of complex numbers is commutative.
Product of four vectors
The product of four vectors can be deduced from the former reasoning. In order
to multiply two pairs of vectors a, b and c, d, rotate the parallelogram formed by a and b
until b' has the direction of c. The product is then the parallelogram formed by a' and d,
but increased by the norm of b and c:
a b c d = a' b' c d = a' b c d = b c a' d
Now let us see the special case where a = c and b = d. If both vectors a, b have the same
direction, the square of their product is a
positive real number:
( a b )2 = a2 b2 > 0

a || b

If both vectors are perpendicular, we must


rotate the parallelogram through a right angle
until b' has the same direction as a (figure
1.9). Then, a' and b are proportional but they
have opposite directions. Therefore, the
square of a product of two orthogonal vectors
is always negative:

Figure 1.9

( a b )2 = a' b' a b = a'bab = a2 b2 < 0

ab

Likewise, the square of an outer product of any two vectors is also negative.
Inverse and quotient of two vectors
The inverse of a vector a is the vector whose multiplication by a gives the unity.
Only vectors that are proportional have a real product. Hence the inverse vector has the
same direction and inverse norm:
a 1 =

a
a

a 1 a = a a 1 = 1

RAMON GONZALEZ CALVET

The quotient of two vectors is the product of one vector by the inverse of the
other vector, which depends on the order of the factors because the product is not
commutative:

a 1 b b a 1
Obviously, the quotient of proportional vectors with the same sense is equal to
the quotient of their norms. When two vectors have different directions, their quotient
can be represented by a parallelogram, which allows extending the concept of vector
proportionality. We say that a is proportional to c as b is to d when their norms are
proportional and the angle between a and c is equal to the angle between b and d1:

a c 1 = b d 1

ac1 =bd1

and

(a, c) = (b, d)

The parallelogram formed by a and b is then similar to that one formed by c and
d. (a, c) is the angle of rotation from the first parallelogram to the second.
The inverse of a product of several vectors is the product of the inverses with an
exchanged order, as you may easily deduce from the associative property:
( a b c ) 1 = c 1 b 1 a 1
Priority of algebraic operations
As in the algebra of real numbers, and in order to simplify the algebraic notation,
we will apply the following priority to the vector operations explained above:
1) Parentheses, whose contents will be first operated.
2) Powers with any exponent (square, inverse, etc.).
3) Outer and inner products, which have the same priority but must be operated
before geometric products.
4) Geometric products.
5) Additions.
As an example, some algebraic expressions are given with the simplified
expression on the left side and its meaning using parentheses on the right side:

abcd=(ab)(cd)
a2 b c + 3 = ( ( a2 ) ( b c ) ) + 3
a+bcde=a+((bc)de)

Sir William Rowan Hamilton defined quaternions as quotients of two vectors in such a way
that similar parallelograms located in the same plane in the three-dimensional space represent
the same quaternion (Elements of Quaternions, posthumously edited in 1866, Chelsea
Publishers 1969, vol. I, see p. 113 and fig. 34). In the vector plane, quaternions are reduced to a
complex numbers. Quaternions were discovered by Hamilton on October 16th, 1843 before
Cliffords geometric product (1878).

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

27

4. TRANSFORMATIONS OF VECTORS
Transformations of vectors are mappings of the vector plane onto itself. Those
transformations preserving the norm of vectors, such as rotations and axial symmetries,
are called isometries and those preserving angles between vectors are said to be
conformal. Besides rotations and axial symmetries, inversions and dilations are also
conformal transformations.

Figure 4.1

Rotations
A rotation through an angle is the
geometric operation consisting of turning a
vector until it forms an angle with the
previous orientation. The positive direction of
angles is counterclockwise (figure 4.1). Under
rotations the norm of any vector is preserved.
According to the definition of the geometric
product, the multiplication of a vector v by a
unit complex number with argument
produces a vector v' rotated through an angle
with respect to v.
v' = v 1 = v ( cos + e12 sin )

This algebraic expression for rotations, when applied to a real or complex number
instead of a vector, modifies its value. However, real numbers are invariant under
rotations and the parallelograms can be turned without changing the complex number they
represent. Therefore this expression for rotations, although it is useful for vectors, is not
valid for complex numbers. In order to remodel it, we factorise the unit complex number
into a product of two complex numbers with half the argument. According to the
permutative property, we can permute the vector and the first complex number whenever
writing the conjugate:
v' = v 1 = v 1/2 1/2 = 1/2 v 1/2 = ( cos

e12 sin

) v ( cos

+ e12 sin

The algebraic expression for rotations now found preserves complex numbers:
z' = 1/2 z 1/2 = z

Let us calculate the rotation of the vector 4e1 through 2 / 3 by multiplying it by the unit
complex number with this argument:
1
2
2
3

= 2 e1 + 2 3 e 2
v' = 4 e1 cos
+ e12 sin
= 4 e1 + e12
3
3
2

2
On the other hand, using the half angle /3 we have:

28

RAMON GONZLEZ CALVET

v' = cos e12 sin 4 e1 cos + e12 sin


3
3
3
3

1
1
3
3
4 e1 + e12
= 2 e1 + 2 3 e 2
= e12

2
2
2
2
Using the expression of half angle, it is not necessary that the complex number have unit
norm because:
z 1 =1/2 z1

z =z1/2

Then, the rotation through an angle can be written as:


v' = z 1 v z
The composition of two successive rotations implies the product of both complex
numbers, whose argument is the addition of the angles of both rotations.
v'' = 1 / 2 v' 1 / 2 = 1 / 2 1 / 2 v 1 / 2 1

/2

= 1 (

+)/2

Axial symmetries

v 1 (+)/2

Figure 4.2

An axial symmetry (also called


reflection) of a vector with respect to a
direction is the geometric transformation that
keeps constant the component with this
direction and reverses the perpendicular
component (figure 4.2). The product of
proportional vectors is commutative and that
of orthogonal vectors is anti-commutative. It is
for this reason that the symmetric vector v'
may be obtained from the multiplication of the
vector v by the unit vector u of the symmetry
axis on the left and right hand sides:
v' = u v u = u ( v|| + v ) u = u v|| u + u v u = v|| v

with u2 = 1

where v|| and v are the components of v respectively proportional and perpendicular to u.
Instead of the unit vector u, any vector d having the axis direction can be
introduced into the expression for axial symmetries, whenever we write its inverse on the
left side of the vector:
v' =

d vd
d

= d 1 v d

Although axial symmetries do not change the absolute value of the angle between
two vectors, they change its sign. Under axial symmetries, real numbers remain invariant

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

29

but complex numbers become conjugate because an axial symmetry generates a


symmetric parallelogram (figure 4.3) and changes the sign of the imaginary part:
z = a + b e12
z' =

a, b real

d ( a + b e12 ) d

d 2 a d 2 b e12
d

Figure 4.3

= a b e12 = z *

For example, let us calculate the axial


symmetry of the vector 3 e1 + 2 e2 with
respect to the direction e1 e2. The
resulting vector will be:
v' =

1
1
( e1 e2 ) ( 3 e1 + 2 e2 ) ( e1 e2 ) = ( e1 e2 ) ( 1 5 e12 ) = 2 e1 3 e2
2
2

Inversions

An inversion of radius r is the geometric transformation that maps every vector v


onto r v 1, that is, onto a vector with the same direction but with a norm equal to r2 / v:
2

v' = r2 v 1

r real

This operation is a generalisation of


the inverse of a vector in geometric algebra
(radius r = 1). It is called inversion of
radius r, because all the vectors with norm
r, whose heads lie on a circle with this
radius, remain unchanged (figure 4.4). The
vectors whose heads are placed inside the
circle of radius r are transformed into
vectors having the head outside and
reciprocally.
Inversion
transforms
complex
numbers into proportional complex
numbers with the same argument (figure 4.5):
v' = r2 v 1

w' = r2 w 1
4

z' = v' w' = r v

Figure 4.4

z=vw

=r vwv

r4 z
z

30

RAMON GONZLEZ CALVET

Dilations

A dilation is the geometric


transformation that enlarges or
shortens a vector, that is, it increases
(or reduces) k times the norm of any
vector
while
preserving
its
orientation. Dilation is simply the
product by a real number k:
v' = k v

Figure 4.5

k real

If k is negative, the vector


direction is reversed.
Most of the transformations
of vectors that will be used in this
book are combinations of these four elementary transformations. Many physical laws are
invariant under some of these transformations. In geometry, from vector transformations
we define transformations of points in the plane, indispensable for solving geometric
problems.
Exercises

4.1 Calculate with geometric algebra what is the composition of an axial symmetry with a
rotation.
4.2 Prove that the composition of two axial symmetries with respect to different directions
is a rotation.
4.3 Consider the transformation under which every vector v multiplied by its transformed
v' is equal to a constant complex z2. Resolve it into elementary transformations.
4.4 Apply a rotation through 2/3 to the vector 3 e1 + 2 e2 and find the resulting vector.
4.5 Find the axial symmetry of the former vector in the direction e1 + e2 .

70

RAMON GONZALEZ CALVET

Perpendicular bisectors and circumcentre

The three perpendicular bisectors of the sides of a triangle meet at a unique point
called circumcentre, the centre of the circumscribed circle. Every point on the
perpendicular bisector of PQ is equidistant from P and Q. Analogously every point on
the perpendicular bisector of PR is equidistant from P and R. The intersection O of both
perpendicular bisectors is simultaneously equidistant from P, Q and R. Therefore O also
belongs to the perpendicular bisector of QR and the three bisectors meet at a unique
point. Since O is equally distant from the three vertices, it is the centre of the
circumscribed circle. Let us use this
condition in order to calculate the equation
Figure 8.3
of the circumcentre:
OP2 = OQ2 = OR2 = d2
where d is the radius of the circumscribed
circle. Using position vectors of each point
we have:
(P O)2 = (Q O)2 = (R O)2
The first equality yields:
P2 2 P O + O2 = Q2 2 Q O + O2
By simplifying and arranging the terms containing O on the left hand side, we have:
2 ( Q P ) O = Q2 P2
2 PQ O = Q2 P2
From the second equality we find an analogous result:
2 QR O = R2 Q2
Now we introduce geometric product instead of inner product into these equations:
PQ O + O PQ = Q2 P2
QR O + O QR = R2 Q2
By subtraction of the second equation multiplied on the right by PQ minus the
first equation multiplied on the left by QR, we obtain:
PQ QR O O PQ QR = PQ R2 PQ Q2 Q2 QR + P2 QR
By using the permutative property on the left hand side and simplifying the right hand
side, we have:
PQ QR O QR PQ O = P2 QR + Q2 RP + R2 PQ

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

71

2 ( PQ QR ) O = P2 QR + Q2 RP + R2 PQ
Finally, multiplication by the inverse of the outer product on the left gives:
O = ( 2 PQ QR ) 1 ( P2 QR + Q2 RP + R2 PQ )
= ( P2 QR + Q2 RP + R2 PQ ) ( 2 PQ QR ) 1
a formula suitable to calculate the coordinates of the circumcentre. For example, let us
calculate the centre of the circle passing through the points P(2, 2), Q(3, 1) and
R(4, 2):
P2 = 8

Q2 = 10

QR = R Q = e1 3 e2

R2 = 20
RP = P R = 2 e1 + 4 e2

PQ = Q P = e1 e2

2 PQ QR = 4 e12
O = ( 8 ( e1 3 e2 ) + 10 ( 2 e1 + 4 e2 ) + 20 ( e1 e2 ) )

e12
4

= e1 2 e2 = ( 1, 2 )
In order to deduce the circle radius, we take vector OP:
OP = P O = P + ( P2 QR + Q2 RP + R2 PQ ) ( 2 PQ QR ) 1
and extract the inverse of the area as common factor:
OP = ( 2 P PQ QR + P2 QR + Q2 RP + R2 PQ ) ( 2 PQ QR ) 1 =
= [ 2 P ( P Q + Q R + R P ) + P2 QR + Q2 RP + R2 PQ ] ( 2 PQ QR ) 1 =
= [ P ( P Q Q P + Q R R Q + R P P R ) + P2 ( R Q ) + Q2 ( P R ) +
+ R2 ( Q P ) ] ( 2 PQ QR ) 1
Simplification gives:
OP = ( P Q R P R Q + P R P P Q P + Q2 P Q2 R + R2 Q R2 P ) ( 2 PQ QR ) 1
= (Q P) (R Q) (P R) ( 2 PQ QR ) 1 = PQ QR RP ( 2 PQ QR ) 1
Analogously:
OQ = QR RP PQ ( 2 PQ QR ) 1

OR = RP PQ QR ( 2 PQ QR ) 1

72

RAMON GONZALEZ CALVET

The radius of the circumscribed circle is the length of any of these vectors:
OP =

PQ

QR

RP

2 PQ QR

PQ
2 sin QRP

QR
2 sin RPQ

RP
2 sin PQR

where we find the law of sines.


Angle bisectors and incentre

The three bisectors of the angles of a


triangle meet at a unique point called incentre.
Any point on the bisector of the angle with
vertex P is equidistant from sides PQ and PR
(figure 8.4). Any point on the angle bisector
passing through Q is also equidistant from sides
QR and QP. Hence their intersection I is
simultaneously equidistant from the three sides,
that is, I is unique, and it is the centre of the
circle inscribed in the triangle.
In order to calculate the equation of the
angle bisector passing through P, we take the
sum of the unit vectors of both adjacent sides:
u=

PQ
PR
+
PQ
PR

v=

Figure 8.4

QP
QR
+
QP
QR

The incentre I is the intersection of the angle bisector passing through P, whose
direction vector is u, and the one passing through Q, with direction vector v:
I=P+ku=Q+mv

k, m real

Arranging terms we find PQ as a linear combination of u and v:


k u m v = Q P = PQ
The coefficient k is:
k=

PQ RP PQ QR
PQ v
=
uv
PQ QR RP + QR RP PQ + RP PQ QR

Since all outer products are equal because they are twice the triangle area, this
expression is simplified:
k=

PQ

RP

RP + PQ + QR

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

73

Then, the centre of the circumscribed circle is:


I =P+ku=P+

PQ
PR

+
PQ + QR + RP PQ
PR
PQ

RP

By taking common denominator and simplifying, we arrive at:


I=

P QR + Q RP + R PQ
QR + RP + PQ

For example, let us calculate the centre of the circle inscribed in the triangle with
vertices:
P(0, 0)

Q(0, 3)

R(4, 0)

PQ = 3

QR = 5

RP = 4

I=

5 (0, 0 ) + 4 (0, 3) + 3 (4, 0) (12, 12 )


=
= (1, 1)
5+4+3
12

In order to find the radius, first we must obtain the segment IP:
IP =

QP RP + RP PQ
QR + RP + PQ

The radius of the inscribed circle is the distance from I to side PQ:
d (I , PQ ) =

IP PQ
PQ

RP PQ
PQ + QR + RP

whence the ratio of radius follows:


radius of circumscribed circle 1 PQ QR RP
=
radius of inscribed circle
2 PQ + QR + RP

Altitudes and orthocentre

The altitude of a side is the segment perpendicular to this side (also called base)
that passes through the opposite vertex. The three altitudes of a triangle intersect at a
unique point called orthocentre. Let us prove this statement by calculating the
intersection H of two altitudes. Since H belongs to the altitude that is perpendicular to
the base QR and passes through vertex P (figure 8.5), its equation is:

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

G=

77

H +2O
3

Hence the centroid is located between the orthocentre and the circumcentre, and its
distance from the orthocentre is double its distance from the circumcentre.
Fermat's theorem

The geometric algebra allows us to prove Fermats theorem in a very easy and
intuitive way.
Over each side of a triangle
ABC we draw an equilateral triangle
Figure 8.7
(figure 8.7). Let T, U and S be the
vertices of the equilateral triangles that
are respectively opposite to A, B and C.
Then, segments AT, BU and CS have
the same length, form angles of 2/3
and intersect at a unique point F, called
the Fermat point. Moreover, the
addition of the three distances from any
point P to each vertex is minimal when
P is the Fermat point, provided that no
interior angle of ABC is greater than
2/3.
First we must demonstrate that BU is obtained from AT by means of a rotation
through 2/3, which will be represented by the complex number t:
AT t = (AC + CT ) t = AC t + CT t

t = cos

2
2
+ e12 sin
3
3

By construction, the vector AC turned through 2/3 is the vector CU, and CT turned
through 2/3 is BC, so that:
AT t = CU + BC = BU

Analogously, one finds CB = BU t and AT = CS t . Therefore, vectors CS, BU and AT


have the same length and each of them is obtained from each other by successive
rotations through 2/3.
Let us see that the sum of distances from P to the three vertices A, B and C is
minimal when P is the Fermat point. First we must prove that the vector sum of PA
turned through 4/3, PB turned through 2/3 and PC is constant and independent of the
point P. (figure 8.8). That is, for any two points P and P it is always true that:
PA t2 + PB t + PC = P'A t2 + P'B t + P'C

A fact that is easily proven by arranging all the terms on one side of the equation:

78

RAMON GONZALEZ CALVET

PP' ( t2 + t + 1 ) = 0

This product is always zero since t2 + t + 1 = 0. Hence, there is a unique point Q such
that:
PA t2 + PB t + PC = QC

Figure 8.8

For any point P, the three segments form a


broken line as shown in figure 8.8. Therefore,
by the triangular inequality we have:
PA + PB + PC QC

When P is the Fermat point F, these segments


form a straight line. Then, the addition of the
distances from F to the three vertices is
minimal provided that no angle of the triangle
is greater than 2/3:
FA + FB + FC = QC PA + PB + PC

Otherwise, some vector among FA t2, FB t, FC has a direction opposite to the others, so
that its length is subtracted from the others and their sum is not minimal.
Exercises

8.1 Napoleons theorem. Over each side of a generic triangle draw an equilateral
triangle. Prove that the centres of these three equilateral triangles also form an
equilateral triangle.
8.2 Leibnizs theorem. Let P be any point in the plane and G the centroid of a triangle
ABC. Then 3 PG2 = PA2 + PB2 + PC2 ( AB2 + BC2 + CA2 ) / 3.
8.3 Apollonius lost theorem. Let A, B and C be three given points in the plane. Every
point G in the plane can then be expressed as a linear combination of these three
points (G is also considered as the centre of masses located at A, B and C with
weights a, b and c 4).
G=aA+bB+cC

with a + b + c = 1

Prove that:
a) a, b, c are the fractions of the area of ABC that are occupied by GBC,
GCA and GAB respectively.
b) The geometric locus of the points P in the plane such that
a PA2 + b PB2 + c PC2 = k is a circle with centre G.

See August Ferdinand Mbius, Der Barycentrische Calcul (1827), p. 17.

122

RAMON GONZALEZ CALVET

hyperbola, both distances are negative, so that the eccentricity is always positive. When
e = 0 both foci are coincident at the centre of a circle. Note that a circle is obtained as the
intersection of a cone with a horizontal plane. In this case, the directrices are the lines at
infinity.
The vectorial equation of a conic is obtained from the polar equation and contains
the radius vector FP. Since FP forms an angle (figure 11.3) with FQ, FP is obtained
from the unit vector of FQ via multiplication by the exponential of e12 and by the norm
of FP yielding:
FP =

1+ e
FQ (cos + e12 sin )
1 + e cos

On the other hand, from the directrix property, one easily finds the following
equation for a conic:
FP2 FT2 = e2 ( FT2 FT FP )2
F, T and e are parameters of the conic, and P(x, y) is the mobile point. Therefore from this
equation we will also obtain a Cartesian equation of second degree. For example, let us
calculate the Cartesian equation of an ellipse with eccentricity , a focus at the point (3,
4) and a vertex at (4,5):
e = 1/2
FT =

F = ( 3, 4 )

Q = ( 4, 5 )

1+ e
FQ = 3 e1 + 3 e 2
e

P = ( x, y )

T = F + FT = ( 6, 7 )

FP = ( x 3 ) e1 + ( y 4 ) e2
The equation of this conic is then:
[ ( x 3 )2 + ( y 4 )2 ] 18 =

1
[ 18 ( 3 (x 3) + 3 ( y 4 ) ) ]2
4

and after simplification it is:


7 x2 2 x y + 7 y2 22 x 38 y + 31 = 0
Chasles theorem

According to this theorem3, the projective cross ratio of any four given points A,
B, C and D on a conic regarded from a point X also lying on this conic is constant,
independently of the choice of the point X (figure 11.7):
{ X, A B C D } = { X', A B C D }
3

Michel Chasles, Trait des sections coniques, Gauthier-Villars, Paris, 1865, p. 3.

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

123

To prove this theorem, let us take into account that points A, B, C, D, X must fulfil
the vectorial equation of the conic. Let us also suppose, without loss of generality, the

Figure 11.7

Figure 11.8

main axis of symmetry having the


direction e1 (this supposition simplifies calculations):
FQ = FQ e1
From now on, , , , , will be the angles that the focal radii FA, FB, FC, FD, FX
form with the main axis with direction vector FQ (figure 11.8). Then:
e cos + e2 sin e1 cos + e2 sin

XA = FA FX = FQ (1 + e ) 1

1 + e cos
1 + e cos

Introducing a common denominator, we find:


XA = FQ

(1 + e ) [e1 (cos cos ) + e2 (sin sin + e sin cos e cos sin )]


(1 + e cos ) (1 + e cos )

From XA and the analogous expression for XC, and after simplification we obtain:
e12 (1 + e ) (sin cos sin cos + sin cos sin cos + sin cos sin cos )
(1 + e cos )(1 + e cos )(1 + e cos )
2

XA XC = FQ 2

= FQ

(1 + e )2 [ sin ( ) + sin ( ) + sin( - ) ]


e12
(1 + e cos ) (1 + e cos ) (1 + e cos )

Using the trigonometric the half-angle identities, the sum is converted into a product of
sines (exercise 6.2):
( 1 + e )2 sin sin sin
2 2 2 e
XA XC = 4 FQ 2
12
( 1 + e cos ) ( 1 + e cos ) ( 1 + e cos )

124

RAMON GONZALEZ CALVET

Likewise, the other outer products are obtained. The projective cross ratio is their
quotient, where the factors containing the eccentricity or the angle are simplified:
XA XC XB XD
{ X , ABCD} =
=
XA XD XB XC

BFD
AFC
sin
2
2 =
2
2
AFD
BFC


sin
sin
sin
sin
2
2
2
2

sin

sin

sin

since is the angle AFC , etc. Therefore, the projective cross ratio of four points A,
B, C and D on a conic is equal to the quotient of the sines of the focal half angles, which
do not depend on X, but only on the positions of A, B, C and D, a fact that is the proof of
Chasles theorem. This statement is trivial for the case of a circle, because the inscribed
angles are half the central angles. However, angles inscribed in a conic vary with the
position of the point X and they differ from half focal angles. In spite of this, the fact that
the quotient of the sines of inscribed angles (projective cross ratio) is equal to the quotient
of the sines of half focal angles is a notable result. For the case of the hyperbola,
remember that the focal radius of a point on the non-focal branch is oriented towards the
focal branch but it has negative norm so that the focal angle is measured with respect to
this orientation.
Tangent and perpendicular to a conic

The vectorial equation of a conic with


the major diameter oriented in the direction
e1 (figure 11.9) is:
FP =

(1 + e )

FQ

1 + e cos

( e1 cos + e2 sin )

The derivation with respect to the angle


gives:

(1 + e ) FQ
d FP
[ e sin + e2 ( e + cos ) ]
=
d
(1 + e cos )2 1

Figure 11.9

This derivative has the direction of the line tangent to the conic at the point P, and its unit
vector t is:
t=

e1 sin + e 2 ( e + cos )
1 + e 2 + 2 e cos

The unit normal vector n is orthogonal to the tangent vector:


n=

e1 ( e + cos ) + e 2 sin
1 + e 2 + 2 e cos

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

147

0
a + b
a + b e1 =

0
a
b

Now it is obvious that there should only be a unique root with odd index, which always
exists for every hyperbolic number:
n

n a + b
a + b e1 =
0

a b

n odd

On the other hand, if a + b > 0 and a b > 0 (right sector) there are four roots with even
index, one in each sector:

n even
n a + b

n a + b

a b

n a + b
right sector
0

0
left sector
n a b

n a + b

0
upper sector
n a b

0
lower sector
a b

If the number does not belong to the right sector, some of the diagonal elements will be
negative and there is no even root. This shows a panorama of the hyperbolic algebra far
from that of complex numbers.

Hyperbolic analytic functions


What conditions should a hyperbolic function f(z) of a hyperbolic variable z
fulfil to be analytic? We want the derivative to be well defined:
f

(z ) = lim f (z + z ) f (z )
z

z 0

that is, this limit must be independent of the direction of z. If f(z) = a + b e1 and the
variable z = x + y e1 , the derivative calculated in the direction z = x is then:

(z ) = a + e1 b
x

while the derivative calculated in the direction z = e1 y becomes:

(z ) = e1 a + b
y

Both expressions must be equal, which results in the conditions of hyperbolic


analyticity:

148

RAMON GONZALEZ CALVET

a b
=
x y

a b
=
y x

and

Note that exponential and logarithm fulfil these conditions and are therefore hyperbolic
analytic functions. More exactly, the exponential function is analytic in the whole plane
while the logarithm function is analytic in the left and right sectors, where the
determinant of the hyperbolic numbers is positive.
By derivation of both identities one finds that analytic functions satisfy the
following hyperbolic partial differential equation, called one-dimensional wave
equation:
2a 2a 2b 2b

=0
x 2 y 2 x 2 y 2
Now we must state the main integral theorem for hyperbolic analytic functions:
if a hyperbolic function is analytic in a certain region in the hyperbolic plane, then its
integral following a closed path C within this region is zero. If the hyperbolic function
is f (z ) = a + b e1 then the integral is:

f (z ) dz = (a + be )(dx + dy e ) = (a dx + b dy ) + e (a dy + b dx )
1

Since C is a closed path, we may apply Greens theorem to write:


b a
a b
= dx dy + e1
dx dy = 0
x y
x y
D
D
where D is the region bounded by the closed path C. Since f(z) fulfils the analyticity
conditions everywhere within D, the integral vanishes.
From here, other theorems analogous to those of complex analysis follow, e. g.:
if f(z) is a hyperbolic analytic function in a simply connected domain D and z1 and z2 are
two points in D then the definite integral:

f (z ) dz
z1

z2

between these points has a unique value independently of the integration path.
Let us see an example. Consider the function f(z) = 1 / ( z 1 ). This function is
only defined if the inverse of z 1 exists, which implies z 10. Of course, this
function is analytic neither at z = 1 nor at the points:
z 12 = 0

( x 1)2 y2 = 0

( x + y 1) ( x y 1) = 0

The lines x + y = 1 and x y = 1 divide the analyticity domain into four sectors. Let us
calculate the integral:

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

149

( 5, 3 )

dz
z 1
( 5 , 3 )

following two different paths in the right sector. The first one is a straight path given by
the parametric equation z = 5 + t e1 (figure 12.2):

(4 t e1 ) dt
dz
dt
e1 =
e1
=

2
z
t
e

1
4
+
t
16

1
( 5 , 3 )
3
3
( 5, 3 )

Owing to symmetry, the integral of the odd


function is zero:
3

4 dt
4+t
e
e = 1 log
=
= e1 log 7
2 1
t
2
4

t
16

3
3
The second path (figure 12.2) is the hyperbola
going from point (5, 3) to (5, 3):

Figure 12.2

(z * 1) dz
(z * 1) dz
dz
=
=

z 1 ( 5, 3) (z 1)(z * 1) ( 5, 3) det (z ) 2 Re (z ) + 1
( 5 , 3 )
( 5, 3 )

( 5, 3 )

( 5, 3 )

Introducing its parametric equation, z = 4 (cosh t + e1 sinh t) we have:


log 2

(4 cosh t 4e1 sinh t 1) (4sinht + 4 e1 sinh t )


16 8 cosh t + 1

log 2

e1 (4 cosh t ) 4 sinh t
dt
17 8 cosh t
log 2
log 2

dt = 4

Due to symmetry, the integral of the hyperbolic sine (an odd function) divided by the
denominator (an even function) is zero. We then split the integral into two integrals and
find its value:
log 2

log 2

15
dt
dt e1
= e1
+ e1
=
2 log 2 17 8 cosh t
2
2
log 2

8 exp(t ) + 2
+ e1 log 2 = e1 log 7
log 8 exp(t ) + 32

log 2
log 2

We now see that the integral following both paths gives the same result, as indicated by
the theorem. In fact, the analytical functions can be integrated directly by using the
indefinite integral:

( 5 , 3)

( 5 , 3)

1
dz
y
( 5, 3)
2
= [ log( z 1) ] ( 5, 3) = log ( x 1) y 2 + e1 arg tanh

z 1
x 1 ( 5 , 3 )
2
( 5 , 3 )
( 5, 3 )

e
x 1+ y
= 1 log
= e1 log 7

x 1 y ( 5 , 3 )
2

164

RAMON GONZALEZ CALVET

Consequently, the eccentricity e of the hyperbola is related to the obliquity of the


transverse plane:
cos = e 2 1 with

1< e < 2

The conjugate diameters of any hyperbola are intersections of this transverse


plane with a pair of orthogonal axial planes; in other words, two radii are conjugate
(figure 11.17) if their projections onto the horizontal plane are the semimajor and
semiminor axes of the prism turned through the same hyperbolic angle :

OQ' = OQ cosh + OS sinh


OS' = OQ sinh + OS cosh
Our Euclidean eyes see these horizontal projections as symmetric lines with respect to
the quadrant bisector. However, they are actually orthogonal because:

OQ' 2 OS' 2 = OQ2 OS2


and, therefore, they can be taken as a new system of orthogonal coordinates. We can
even draw a new picture with the new diameters on the Cartesian axis.
The central equation of the hyperbola using the rotated axes is:
OP = (OQ' cosh( ) + OS' sinh( ))

which shows that a hyperbolic rotation of the coordinate axes has been made with respect
to the principal diameters of the hyperbola.
The law of sines and cosines

Since the norm of the area is identical in both the Euclidean and the hyperbolic
planes, a parallelogram is divided by its diagonal into two triangles with equal area.
This statement is somewhat subtle since the Euclidean congruence of triangles is not
valid in the hyperbolic plane. We will return to this question later. Now we only need to
know that the area of a hyperbolic triangle is half the outer product of any two sides.
Following the perimeter of a triangle, let a, b, and c be its sides respectively
opposite to the angles , and . Then, the angles formed by the oriented sides and the
angles , and are supplementary:
a b=bc=ca

a
sinh

b
sinh

a b sinh = b c sinh = c a sinh

c
sinh

which is the law of sines.


Since a triangle is a closed polygon, a + b + c = 0 and we have:

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

a 2 = ( b c ) = b 2 + c 2 + 2 b c
2

165

a 2 = b 2 + c 2 2 b c cosh

which is the law of cosines. And also:


b 2 = a 2 + c 2 2 a c cosh

Figure 13.8

c 2 = a 2 + b 2 2 a b cosh

When applying both theorems, we must


be careful with the sides having
imaginary length and the signs of the
angles and trigonometric functions.
As an application of the law of
sines and cosines, consider the
hyperbolic triangle having the vertices
A(5, 3), B(1, 0), C(10, 1), whose sides
have real norm (figure 13.8):

( 4 )2 ( 3)2

AB = B A = 4 e 2 3 e 21

AB =

BC = C B = 9 e 2 + e 21

BC = 9 2 12 = 80

CA = A C = 5 e 2 + 2 e 21

CA =

( 5)2 2 2

= 7

= 21

BC 2 = CA 2 + AB 2 2 CA AB cosh

cosh =

CA 2 = AB 2 + BC 2 2 AB BC cosh

cosh =

AB 2 = BC 2 + CA 2 2 BC CA cosh

cosh =

26
7 3
33

4 35
47
4 105

From where it follows that:

= 1.3966... e12

= 0.8614...

= 0.5352...

I have obtained their signs from the definition of the angles =BAC, =CBA,
=ACB and the geometric plot (figure 13.8). Note that + + = e12 and that
they fulfil the law of sines:

BC
sinh

CA
sinh

AB
sinh

170

RAMON GONZALEZ CALVET

FOURTH PART: PLANE PROJECTIONS OF THREE-DIMENSIONAL SPACES


The complete study of the geometric algebra of the three-dimensional spaces falls
out of the scope of this book. However, due to the importance of Earth charts and of
Lobachevskys geometry -the first one is more practical and the second one more
theoretical-, I have written this last section. In order to make the explanations clearer, the
three-dimensional geometric algebra has been reduced to the minimal concepts,
enhancing the plane projections.
The geometric quality of being Euclidean or pseudo-Euclidean is not the signature
+ or of a coordinate, but the fact that two coordinates have the same or different
signature, in other words, it is a characteristic of the plane. For instance, a plane with
signatures + + is equivalent, from a geometric point of view, to another with .
Therefore, only two kinds of three-dimensional spaces exist: the room space where all the
planes are Euclidean (signatures + + + or ), and the pseudo-Euclidean space, which
has one Euclidean plane and two orthogonal pseudo-Euclidean planes (signatures + or
+ + ).
14. SPHERICAL GEOMETRY IN THE EUCLIDEAN SPACE
The geometric algebra of the Euclidean space
A vector of the Euclidean space is an oriented segment in this space with direction
and sense that may represent other physical magnitudes such as forces, velocities, electric
fields, etc. The set of all segments (geometric vectors) together with their addition
(parallelogram rule) and their product by real numbers (dilation of vectors) has a structure
of vector space, symbolised here by V3. Every vector in V3 has the form:
v = v1 e1 + v 2 e 2 + v 3 e 3
where ei are three unit perpendicular vectors, which are a basis of the space. If we define
an associative product (geometric or Clifford product) as a generalisation of the
multiplication of vectors in the Euclidean plane, we will arrive at:

ei2 = 1

and

ei e j = e j ei for i j

In general, the square of a vector is the square of its norm and perpendicular vectors
anticommute whereas proportional vectors commute.
The geometric algebra generated by the space V3 has eight dimensions:
Cl (V3 ) = Cl 3, 0 = 1, e1 , e 2 , e 3 , e 23 , e 31 , e12 , e123

Let us see the product of two vectors in more detail:


v w = (v1 e1 + v 2 e 2 + v 3 e 3 )(w1 e1 + w2 e 2 + w3 e 3 ) = v1 w1 + v 2 w2 + v 3 w3
+ (v 2 w3 v3 w2 ) e23 + (v3 w1 v1 w3 ) e31 + (v1 w2 v 2 w1 ) e12

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

171

The product (or quotient) of two vectors is called a quaternion1. Quaternions are the even
subalgebra of Cl3, 0 that generalise complex numbers to the three-dimensional space.
Splitting a quaternion into the real and bivector parts, we obtain the inner (or scalar)
product and the outer (or exterior) product respectively:
v w = v1 w1 + v 2 w2 + v 3 w3
v w = (v 2 w3 v 3 w2 ) e 23 + (v 3 w1 v1 w3 ) e 31 + (v1 w2 v 2 w1 ) e12
Bivectors are oriented plane surfaces and indicate the direction of planes in space.
Those who are acquainted with vector analysis will say that both vectors and bivectors are
the same thing. This confusion was originated by Hamilton2 himself, and continued by the
founders of vector analysis, Gibbs and Heaviside. However, vectors and bivectors are
different things just as physicists have experienced and know since a long time ago. The
proper vectors are usually called polar vectors while the pseudo-vectors, which are
actually bivectors, are usually called axial vectors. The following magnitudes are
vectors: of course a geometric segment, but also a velocity, an electric field, the
momentum, etc. On the other hand, the oriented area is, of course, a bivector, but also the
angular momentum, the angular velocity and the magnetic field. A criterion to distinguish
both kinds of magnitudes is the reversal of coordinates, which changes the sense of
vectors while leaves bivectors unchanged.
The product of two bivectors yields a real number plus a bivector. Both parts can
be separated as symmetric and antisymmetric product. The symmetric product is a real
number and its negative value will be denoted here by the symbol while the
antisymmetric product is also a bivector and will be denoted here by the symbol :
vw=

1
(v w + w v ) = v 23 w23 + v 31 w31 + v12 w12
2

vw=

1
(v w w v )
2

= (v 31 w12 v12 w31 ) e 23 + (v12 w23 v 23 w12 ) e31 + (v 23 w31 v 31 w23 ) e12

v w = v w v w
Let us see what happens with the outer product of three vectors. According to the
extension theory of Grassmann, the product u v w is the oriented volume obtained
from the surface the bivector u v represents by a parallel translation along the
segment w:

From this definition, Hamilton deduced the multiplication rule of i, j, k. I recommend the
reading of the initial chapters of the Elements of Quaternions, especially the section 2 First
Motive for naming the Quotient of two Vectors a Quaternion in chapter I, p. 110.
2
This confusion is due to the fact that vectors and bivectors are dual spaces of the algebra Cl3, 0.
However, duality between vectors and bivectors does not exist at higher dimensions, although
there are also dualities among other spaces.

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

189

15. HYPERBOLOIDAL GEOMETRY IN THE PSEUDO-EUCLIDEAN SPACE


The geometric algebra of the pseudo-Euclidean space
A vector in the pseudo-Euclidean space is an oriented segment in this space with
direction and sense. The set of all segments (vectors) together with their addition
(parallelogram rule) and the product by real numbers (dilation of vectors) has a structure
of vector space, symbolised here by W3. Every vector in W3 has the form:
v = v1 e1 + v 2 e 2 + v 3 e 3
where ei are three unit perpendicular vectors, which constitute the basis of the space.
The square of the norm of a vector is:

= v12 + v 22 v32

which determines the geometric properties of this space, very different from the
Euclidean space. Now we define an associative product (geometric or Clifford product)
as a generalisation of those multiplications of vectors defined for the Euclidean and
hyperbolic planes. Imposing the condition that the square of the norm must be equal to
the square of the vector, we find:

= v2

e12 = e22 = 1

e32 = 1

and

ei e j = e j ei for i j

From the basis vectors one deduces that the geometric algebra generated by the space
W3 has eight components:
Cl (W3 ) = Cl 2, 1 = 1, e1 , e2 , e3 , e23 , e31 , e12 , e123

Let us see with more detail the product of two vectors:


v w = (v1 e1 + v 2 e2 + v3 e3 ) (w1 e1 + w2 e2 + w3 e3 ) = v1 w1 + v 2 w2 v3 w3
+ (v 2 w3 v3 w2 ) e23 + (v3 w1 v1 w3 ) e31 + (v1 w2 v 2 w1 ) e12
I shall call the product (or quotient) of two vectors in W3 a tetranion. Tetranions are the
even subalgebra of Cl2,1, which generalises complex and hyperbolic numbers to the
pseudo-Euclidean space. Let t* be the conjugate of a tetranion t.
t = a + b e23 + c e31 + d e12
Then, the norm of t is given by:
t = t t * = a2 b2 c2 + d 2

t* = a b e23 c e31 d e12

190

RAMON GONZALEZ CALVET

2
2
because e23
= e31
= 1 and e122 = 1 . Consequently, e12 = 1 , according to the fact that it

represents a Euclidean plane whereas e23 , e31 have imaginary norm.


Splitting the tetranion product of two vectors into the real and bivector parts, we
obtain the inner (or scalar) product and the outer (or exterior) product respectively:
v w = v1 w1 + v 2 w2 v3 w3
v w = (v 2 w3 v 3 w2 ) e 23 + (v 3 w1 v1 w3 ) e 31 + (v1 w2 v 2 w1 ) e12
Here bivectors are also oriented plane surfaces indicating the direction of planes in the
pseudo-Euclidean space. As before, vectors and bivectors are different things. Physics
have also experienced this fact: in Minkowskis space, the electromagnetic field is a
bivector whereas the tetrapotential is a vector. On the other hand, oriented areas are, of
course, bivectors. As a criterion to distinguish both kinds of magnitudes one also uses
the reversal of coordinates, which changes the sense of vectors while leaving invariant
bivectors.
Two vectors are said to be orthogonal if their inner product vanishes:

vw vw=0
Thus, the outer product is the product by the orthogonal component and the inner
product is the product by the proportional component:
v w = v w||

v w = v w

The product of two bivectors yields a tetranion. The real and bivector parts can
be separated as symmetric and antisymmetric products. The symmetric product is a real
number whose negative value will be denoted here by the symbol , whereas the
antisymmetric product with negative sign, denoted here by the symbol , is also a
bivector:
vw =

1
(v w + w v ) = v23 w23 v31 w31 + v12 w12
2

vw =

1
(v w w v )
2

= (v31 w12 v12 w31 ) e23 + (v12 w23 v 23 w12 ) e31 (v 23 w31 v31 w23 ) e12

v w = v w v w
The outer product of three vectors has the same expression as for Euclidean
geometry and this is a natural outcome of the extension theory: the product u v w is
the oriented volume generated by the surface represented by the bivector u v when it
is translated parallelly along the segment w:

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

u1

v1

w1

u v w = u2
u3

v2
v3

w2 e123
w3

Since e123 = e12

191

e3 , the pseudoscalar e123 has imaginary norm.

Finally, let us see how the product of three vectors u, v and w is. The vector v
can be resolved into a component coplanar with u and v and another component
perpendicular to the plane u-v:

u v w = u v || w + u v w
Let us analyse the permutative property now. For both Euclidean and hyperbolic planes
we found u v w w v u = 0. For the pseudo-Euclidean space the permutative property
becomes:
u v w w v u = u v w w v u = v ( u w + w u )
= 2 v u w = 2 v u w = 2 u v w

I take the same algebraic priorities as in the former chapter: all the abridged
products must be operated before the geometric product, which is a convention adequate
to the fact that the abridged products have to be developed as sums of geometric
products.
The hyperboloid of two sheets (Lobachevskian surface)

According to Hilbert (Grundlagen der Geometrie, Anhang V) the complete


Lobachevskys plane cannot be represented by a smooth surface with a constant
curvature as proposed by Beltrami. However,
this result only concerns surfaces in the
Euclidean space. The surface whose points are
placed at a fixed distance from the origin in a
pseudo-Euclidean space (the two-sheeted
hyperboloid) is the surface sought by Hilbert that
realises Lobachevskys geometry1. It is known
that it has a characteristic distance like the radius
of the sphere. Since all the spheres are similar,
we only needed to study the unit sphere.
Figure 15.1
Likewise, all the hyperboloidal surfaces
2
2
2
2
x +y z =r
are similar and the
hyperboloid with imaginary unit radius (figure
15.1):
z2 x2 y2 =1
1

The reader will find a complete study of the hyperboloidal surface in Faber, Foundations of
Euclidean and Non-Euclidean Geometry, chap. VII. The Weierstrass model.

206

RAMON GONZALEZ CALVET

tanh
2
= tanh 0

0
tanh
2

About the congruence of geodesic triangles


Two geodesic triangles on the hyperboloid having the same angles also have the
same sides and are said to be congruent. This follows immediately from the rotations in
the pseudo-Euclidean space, which are expressed by means of tetranions like
quaternions are used for the rotations of Euclidean space. However, this falls out of the
scope of this book and will not be treated. The reader should perceive, in spite of his
Euclidean eyes9, that all the points on the hyperboloid are equivalent because the
curvature is always the imaginary unit and the surface is always perpendicular to the
radius. Thus the pole (vertex of the hyperboloid) is no special point, and any other point
may be chosen as a new pole provided that the new axis are obtained from the old ones
under a hyperbolic rotation.
The hyperboloid of one sheet
Whereas the Lobachevskian surface has imaginary unit radius, the one-sheeted
hyperboloid (figure 15.9) has real unit radius and its equation is:
x2 + y2 z 2 = 1
Under the central projection, the upper
part of the hyperboloid is projected outside
the Beltrami disk. However, we must take
into account the fact that, if we want to
represent the whole surface, each point in
the projection plane has duplicity and
represents two points: one on the upper
part and the other on the lower part of the
hyperboloid. This fact has significance and
must be taken into account because the
one-sheeted hyperboloid is a continuous
surface. Something similar happens for the
Figure 15.9
two-sheeted hyperboloid, but both sheets
are not connected so that the
Lobachevskian surface is only one of the sheets and projections apply to it.
Points lying in the x-y plane, which separate the upper and lower parts of the
one-sheeted hyperboloid, are projected onto the line at infinity in the plane z = 1 while
9

In fact, eyes are not Euclidean but a very perfect camera where images are projected onto a
spherical surface. The principles of projection are likewise applicable to the pseudo-Euclidean
space. Our mind, accustomed to the ordinary space (we learn its Euclidean properties in the first
years of our life), deceives us when we wish to perceive the pseudo-Euclidean space.

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

207

points at infinity on the hyperboloid are projected (as for Lobachevskys surface) onto
the limit circle of the Beltrami disk. Geodesics are intersections of the hyperboloid with
central planes10 z = a x + b y. There are three kinds of geodesics depending on the
orientation of the plane: hyperbolas if a2 + b2>1, ellipses if a2 + b2<1, straight lines and
at the same time generatrices of the hyperboloid if a2 + b2=1. The central projection
maps every geodesic onto a line that either cuts the limit circle (hyperbola) or does not
intersect it (ellipse), or is tangent to the limit circle (generatrix of the hyperboloid).
The three former cases yield three different kinds of arc length: For hyperbolas
the square of the arc length is negative; generatrices have null arc length; ellipses have
real arc length such as Lobachevskian metric. Thus, the tangent plane to any point on
the one-sheeted hyperboloid is a hyperbolic plane, in contrast with the fact that the
tangent plane to any point on the Lobachevskys surface is Euclidean.
Central projection and arc length on the one-sheeted hyperboloid

Since geodesics are projected onto straight lines under the central projection, it is
the suitable projection to calculate arc lengths. The deduction steps resemble those for
the two-sheeted hyperboloid although some signs change. As before:
u=

x
z

v=

y
z

From the hyperboloid equation x 2 + y 2 z 2 = 1 we obtain:


x=

y=

u2 + v2 1

z=

u 2 + v2 1

1
u 2 + v2 1

The differential of the arc length on the hyperboloid is:


ds = dx e1 + dy e 2 + dz e 3
ds = dx + dy dz
2

(1 v )du
=
2

+ 2 u v du dv + 1 u 2 dv 2

(u

+ v 1
2

where the only difference with regard to Lobachevskian metric is a minus sign.
Geodesics are intersections of central planes with the hyperboloid and are
projected onto straight lines under the central projection. So we take:
v = k u+l

Substitution into the arc length differential gives:

10

This proof has been left as exercise 15.7.

274

RAMON GONZALEZ CALVET

12.4 From the decomposition theorem we have:


sin ( x + y e1 ) =

1 + e1
1 e1
sin ( x + y ) +
sin ( x y ) = sin x cos y + e1 cos x sin y
2
2

12.5 From the analogous of de Moivres identity we have:

(cosh + e1 sinh )4 cosh 4 + e1 sinh 4


cosh 4 cosh 4 + 6 cosh 2 sinh 2 + sinh 4
sinh 4 4 cosh 3 sinh + 4 cosh sinh 3
12.6 The analytical continuation of the real logarithm is:
log( x + y e1 ) =

1 + e1
1 e1
log( x + y ) +
log( x y )
2
2

e
1
x+ y
log(x 2 y 2 ) + 1 log
2
2
xy

It may be rewritten in the form:


= log x 2 y 2 + e1 arg tanh

y
x

12.7 For the straight path z = t e1 we have:


e1

2e
t 3
z
dz
e
t
dt
e
= 1
=
=
1
1

e
3
3 1
1
1
1

For the circular path z = cos t + e1 sin t we have:


e1

2
z dz =

e1

/2

(cos t + e

sin t ) ( sin t + e1 cos t ) dt


2

/ 2

/2

2e
2
2

= cos t + cos 3 t + e1 sint sin 3 t


= 1
3
3
3

/ 2

Using the indefinite integral, we find the same result:


e1

e1

z3
2 e1
=
=
z
dz

3
3 e1
e1
2

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

275

12.8 The proof is analogous to the exercise 3.12: turn the hyperbolic numbers df, dz into
hyperbolic vectors by multiplying them by e2 on the left.
12.9 Let us prove by induction that any power with positive exponent fulfils the
decomposition theorem:
x + y e1 =

1 + e1
(x + y ) + 1 e1 (x y )
2
2

(x + y e1 )n+1 = (x + y e1 )n (x + y e1 )
1 + e1
(x + y )n + 1 e1 (x y )n
=
2
2

1 e1
1 + e1

2 ( x + y ) + 2 ( x y )

1 + e1
(x + y )n+1 + 1 e1 (x y )n+1
2
2

n N

We have already seen that:


1 + e1 1
1 e1 1
1
=
+
x + y e1
2 x+ y
2 x y

The fact that the powers with negative exponents also fulfil the decomposition theorem
is also proven by induction in the same way as above:
1

(x + y e1 )

1 + e1
1 e1
1
1
+
n
2 (x + y )
2 ( x y )n

n N

Therefore, the Laurent series also fulfils the decomposition theorem:

a (x + y e )

n =

1 + e1
2

a (x + y )

n =

1 e1
n
a n (x y )

2 n =

and we have:
f ( x + y e1 ) =

1 + e1
1 e1
f (x + y ) +
f (x y )
2
2

13. The hyperbolic or pseudo-Euclidean plane


13.1 If the vertices of the triangle are A(2, 2), B(1, 0) and C(5, 3) then:

AB = (1, 2)

BC = (4, 3)

CA = (3, 1)

nZ

TREATISE OF PLANE GEOMETRY THROUGH GEOMETRIC ALGEBRA

305

CHRONOLOGY OF THE GEOMETRIC ALGEBRA

1679 Letters from Leibniz to Huygens on the characteristica geometrica.


1799 Publication of Om Directionens analytiske Betegning by Caspar Wessel with scarce
diffusion.
1805 Birth of William Rowan Hamilton in Dublin.
1806 Publication of Essai sur une manire de reprsenter les quantits imaginaires dans
les constructions gomtriques by Jean Robert Argand.
1809 Birth of Hermann Gnther Grassmann in Stettin.
1818 Death of Wessel
1822 Death of Argand.
1827 Publication of Der barycentrische Calcul by Mbius in Leipzig.
1831 Birth of James Clerk Maxwell in Edinburgh.
1831 Birth of Peter Guthrie Tait.
1839 Birth of Josiah Willard Gibbs in New Haven.
1843 Discovery of the quaternions by Hamilton.
1844 Publication of the first edition of Die Lineale Ausdehnungslehre where Grassmann
presents the anticommutative product of geometric unities (outer product).
1845 Birth of William Kingdon Clifford in Exeter.
1847 Publication of Geometric Analysis with a foreword by Mbius, memoir with which
Grassmann won the prize that had been offered to whom could develop Leibnizs
characteristica geometrica.
1850 Birth of Oliver Heaviside in London.
1853 Publication of Lectures on Quaternions where Hamilton introduces the nabla
(gradient) operator.
1862 Publication of the second edition of Die Ausdehnungslehre.
1864 Publication of A dynamical theory of the electromagnetic field by Maxwell, where
he defines the divergence and the curl.
1865 Death of Hamilton.
1866 Posthumous publication of Hamiltons Elements of Quaternions.
1867 Publication of Elementary Treatise on Quaternions by Tait.
1873 Publication of Introduction to Quaternions by Kelland and Tait.
1873 Maxwell publishes the Treatise on Electricity and Magnetism where he writes the
equations of electromagnetic field with quaternions.
1877 Publication of Grassmanns paper Der Ort der Hamiltonschen Quaternionen in der
Ausdehnungslehre.
1877 Death of Grassmann.
1878 Publication of the paper Applications of Grassmann's Extensive Algebra by
Clifford where he makes the synthesis of the systems of Grassmann and Hamilton.
1879 Death of Maxwell.
1879 Death of Clifford.
1880 Publication of Lipschitzs Principes dun calcul algbrique qui contient comme
espces particulires le calcul des quantits imaginaires et des quaternions.
1881 Private printing of Elements of Vector Analysis by Gibbs.
1886 Publication of Lipschitz Untersuchungen ber die Summen von Quadraten.
1886 Publication of Gibbs paper On multiple algebra.
1888 Publication of Peanos Calcolo geometrico secondo lAusdehnungslehre di H.
Grassmann preceduto dalle operazione della logica deduttiva.

306

RAMON GONZALEZ CALVET

1891 Oliver Heaviside publishes The elements of vectorial algebra and analysis in The
Electrician Series.
1895 Publication of Peanos Saggio di Calcolo Geometrico.
1901 Death of Tait.
1901 Wilson publishes Gibbs lessons in Vector Analysis.
1903 Death of Gibbs.
1925 Death of Heaviside.
1926 Wolfgang Pauli introduces his matrices to explain the electronic spin.
1928 Publication of the paper The Quantum Theory of Electron, where Paul A. M.
Dirac defines a set of 44 anticommutative matrices built from Paulis matrices.

This comparative diagram of life and works of the authors of (or related with) the
geometric algebra visualises and summarises the chronology. The XIX century may be
properly called the century of the geometric algebra. Note the premature death of
Clifford, which caused the delay in the development of the geometric algebra throughout
the XX century.

1800
|

1820
|

1840
|

1860
|

1880
|

1900
|

1920
|

Wessel*********************
Argand***********************
Hamilton
*******************************
Grassmann
***********************************
Maxwell
*************************
Tait
************************************
Gibbs
*********************************
Clifford
******************
Heaviside
**************************************
Om Directionens
|
Essai sur ...
|
Der barycentrische Calcul
|
Die Ausdehnungslehre (1st ed.)
|
Lectures on Quaternions
|
Die Ausdehnungslehre (2nd ed.)
|
Elements of Quaternions
|
Elementary Treatise on Quaternions
|
Treatise on Electricity and Magnetism
|
Der Ort der Hamiltonschen Quat.in der Ausdehnungslehre |
Applications of Grassmann's Extensive Algebra
|
Elements of Vector Analysis
|
The Elements of Vectorial Algebra and Analysis
Vector Analysis

|
|

Das könnte Ihnen auch gefallen